Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Received: 25 August 2022

| Accepted: 12 July 2023

DOI: 10.1111/jpy.13373

RESEARCH ARTICLE

Gene-­rich plastid genomes of two parasitic red algal


species, Laurencia australis and L. verruciformis
(Rhodomelaceae, Ceramiales), and a taxonomic revision
of Janczewskia

Maren Preuss1,2 | Pilar Díaz-­Tapia3,4 | Heroen Verbruggen5 |


Giuseppe C. Zuccarello2

1
National Institute of Water and
Atmosphere Research, Wellington, Abstract
New Zealand Parasitic red algae are an interesting system for investigating the genetic
2
School of Biological Sciences, Victoria changes that occur in parasites. These parasites have evolved independently
University of Wellington, Wellington, New
Zealand multiple times within the red algae. The functional loss of plastid genomes
3
Coastal Biology Research Group, Faculty can be investigated in these multiple independent examples, and fine-­scale
of Sciences and Centre for Advanced patterns may be discerned. The only plastid genomes from red algal parasites
Scientific Research, University of A known so far are highly reduced and missing almost all photosynthetic genes.
Coruña, A Coruña, Spain
4
Our study assembled and annotated plastid genomes from the parasites
Instituto Español de Oceanografía
(IEO-­CSIC), Centro Oceanográfico de A Janczewskia tasmanica and its two Laurencia host species (Laurencia elata
Coruña, A Coruña, Spain and one unidentified Laurencia sp. A25) from Australia and Janczewskia ver-
5
School of BioSciences, University of ruciformis, its host species (Laurencia catarinensis), and the closest known
Melbourne, Parkville, Victoria, Australia free-­living relative (Laurencia obtusa) from the Canary Islands (Spain). For
Correspondence the first time we show parasitic red algal plastid genomes that are similar in
Maren Preuss, National Institute of size and gene content to free-­living host species without any gene loss or
Water and Atmosphere Research, genome reduction. The only exception was two pseudogenes (moeB and
Wellington 6021, New Zealand; School of
Biological Sciences, Victoria University of ycf46) found in the plastid genome of both isolates of J. tasmanica, indicating
Wellington, PO Box 600, Wellington 6140, potential for future loss of these genes. Further comparative analyses with the
New Zealand. three highly reduced plastid genomes showed possible gene loss patterns, in
Email: maren.preuss@niwa.co.nz
which photosynthetic gene categories were lost followed by other gene cat-
Funding information egories. Phylogenetic analyses did not confirm monophyly of Janczewskia,
Australian Biological Resources and the genus was subsumed into Laurencia. Further investigations will de-
Study, Grant/Award Number: Activity
4-­G046WSD; Royal Society Te Apārangi, termine if any convergent small-­scale patterns of gene loss exist in parasitic
Grant/Award Number: Marsden Fast-­ red algae and how these are applicable to other parasitic systems.
Start/ 19-­VUW-­0 06; Xunta de Galicia,
Grant/Award Number: ED481D/ 2017/011/
03IN858A2019-­1630129 KEYWORDS
Janczewskia, Laurencia, moeB, obligate hemiparasites, parasite origin, phylogenomics,
Editor: M. Oliveira pseudogenes, red algal parasites, reductive plastid evolution, ycf46

I N TROD UC T I O N in nearly all organisms with photosynthetic ancestors,


and besides photosynthetic capability, these organ-
Comparative plastid genome studies in parasitic sys- elles provide many other functions including amino
tems are important for understanding evolutionary acid, fatty acid, and lipid synthesis (Wise, 2007). In
changes during the transition from a free-­living to a parasites, plastid genomes commonly lose genes,
parasitic life strategy. Plastid genomes are observed and in many cases, highly reduced plastids have

Abbreviations: CTAB, cetyltrimethylammonium bromide; HTS, high-­throughput sequencing.

950 | © 2023 Phycological Society of America. wileyonlinelibrary.com/journal/jpy Journal of Phycology. 2023;59:950–962.


15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 951

lost their photosynthetic capability (e.g., de Koning & grow on Laurencia spp. and occasionally on closely re-
Keeling, 2006; Kayama et al., 2020). Functional and lated genera, such as Chondria and Osmundea (Pre-
physical reductions in plastid genomes occur over uss et al., 2017). Molecular data (based on cox1, nrLSU,
time, and the degree of degradation in parasitic plastid and nrSSU genes) have not support the generic status
genomes reflects the transition from autotrophy to par- of Janczewskia, as it is not monophyletic, and the two
asitism, where organisms with intact plastid genomes investigated parasite species evolved independently
are less reliant on their hosts (for nutrition) than par- within the genus Laurencia with J. hawaiiana distantly
asites with highly reduced plastid genomes (Banerjee related to its host L. mcdermidiae, whereas J. morim-
et al., 2022; Wicke & Naumann, 2018). The early stages otoi is nested within its host L. nipponica (Kurihara
of this transition are of interest for identifying any pat- et al., 2010). No other Janczewskia species have been
terns of gene loss that may be common among para- investigated genetically, and there are no genomic data
sites. Fine-­scale patterns in recently diverged parasites available.
are hard to investigate, as few study systems provide: For this study, we investigated the pigmented red
(i) multiple independently evolved parasites in order to algal parasites Janczewskia tasmanica from Australia
observe different stages of gene and functional loss and J. verruciformis from the Canary Islands, as prelim-
and (ii) comparisons with closely related free-­ living inary molecular data showed the presence of rbcL in
organisms. these parasites, indicating that photosynthetic capacity
Parasitic red algae are excellent systems for investi- may have been retained in contrast to the other three
gating these transitions, as these parasites evolved in- known circular-­ mapping parasite plastid genomes
dependently multiple times. These parasites are highly (Preuss et al., 2020; Salomaki et al., 2015; Salomaki &
diverse with almost 130 species currently described Lane, 2019). The overall aims of this study were to: (i)
(Freese & Lane, 2021; Preuss et al., 2017; Preuss & compare plastid genome architecture and characteris-
Zuccarello, 2018; Preuss & Zuccarello, 2019). Molecular tics between parasites and their hosts and between all
phylogenetics show different evolutionary relationships parasites and (ii) investigate the evolutionary origin of
of these parasites to their hosts (e.g., Goff et al., 1996; Janczewskia.
Kurihara et al., 2010; Preuss & Zuccarello, 2018). Most
parasites are closely related to their hosts, often group-
ing with the host as its closest relative, indicating that M ATER IAL S AN D M E T H ODS
the parasite evolved recently (Goff et al., 1997; Ng
et al., 2014). In other cases, the parasites are phylo- Dried silica gel samples of two dark reddish pigmented
genetically distant from their hosts, and their closest Janczewskia parasites and their associated hosts were
relative is a non-­host species (Goff et al., 1996; Kuri- collected in Australia and the Canary Islands (Spain);
hara et al., 2010; Preuss & Zuccarello, 2018; Zuccarello samples were also collected of one of the parasites
et al., 2004). closest free-­living relatives (Table S1 in the Supporting
The discovery of the highly reduced plastid genome Information). DNA extraction of all samples followed a
in the unpigmented parasite Choreocolax polysipho- modified cetyltrimethylammonium bromide protocol
niae (Salomaki et al., 2015) demonstrated that parasitic (Zuccarello & Lokhorst, 2005) and was either used for
red algae can retain their own plastid genome instead high-­throughput sequencing (HTS) or for Sanger se-
of only maintaining the acquired plastid genome of their quencing to amplify individual genes.
host (Goff & Coleman, 1995). Shortly after that study, For HTS, samples from the Canary Islands were
two more highly reduced plastid genomes were identi- individually sequenced (n = 1), whereas samples from
fied in Harveyella mirabilis (Salomaki & Lane, 2019) and Australia were pooled by combining each host (n = 1)
Pterocladiophila hemisphaerica (Preuss et al., 2020). and parasite (n = 1–­ 5). Multiple parasites were com-
Comparative plastid genomic analysis showed the loss bined when possible due to their size. In addition, the
of almost all photosynthetic genes (with the exception non-­host species Laurencia obtusa was sequenced,
of petF) and a conserved core of housekeeping genes. as our preliminary results showed it to be a close rela-
To date no further full plastid genomes of parasitic red tive of Janczewskia verruciformis. Libraries were con-
algae have been sequenced. Further studies will be im- structed using Illumina TruSeq Nano DNA Library Prep
portant for revealing the extent of the changes seen in kits and were sequenced on a NovaSeq 6000 by Mac-
parasite plastid genomes and if there are any shared rogen Inc. (Seoul, Korea) with 150 bp paired-­end reads
patterns that can be linked to the change from free-­ and resulting in between 22,529,862 and 23,207,636
living to parasite. total read counts of raw data reads. Assembly and an-
The parasitic genus Janczewskia was described notation of the plastid genomes followed as previously
based on J. verruciformis growing on Laurencia ob- described in Preuss et al. (2020). Briefly, sequenced
tusa in Naples, Italy (Guiry & Guiry, 2022; Solms-­ reads were processed in two separate workflows (i)
Laubach, 1877). Currently, a total of 11 Janczewskia trimming and assembly with the CLC Genomic Work-
species are taxonomically accepted, and these mostly bench 20.0.05 (CLC Bio, Aarhus, Denmark) and (ii)
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
952 |    PREUSS et al.

trimming in trimmomatic (Bolger et al., 2014) and then

508,866

1,263,035

799,726

2,949,102
501,479

1,601,509
assembly in SPAdes 3.14.0 (Prjibelski et al., 2020).
reads
Total Candidate contigs were identified using a custom data-
base containing known red algal genes, and annotation
was carried out in MFannot (Beck & Lang, 2010; https://
coverage

megas ​ u n .bch.umont ​ r eal.ca/cgi- ­b in/mfann ​ o t/mfann​


Mean

2585

1388
1091
otInt​erface.pl) and Aragorn (Laslett & Canback, 2004;

694
442

428
(X)

http://www.ansik​te.se/ARAGO​RN/) after circularity was


checked. Missing gene annotations identified through
(domain
Intron II

open reading frames (ORFs) and comparisons of anno-


tations between different species in MAUVE alignments
V)

1
1

1
were manually added in Geneious Prime 2021.2.2
Plastid genome characteristics of the Janczewskia parasites and their associated host species, plus one closely related non-­host species.

ncRNA

(https://www.genei​ous.com/). Genes were annotated


as pseudogenes if there were frameshift mutations due
1

1
1

1
to insertion/deletions or mutations leading to stop co-
dons. Frame shift mutations in identified pseudogenes
rRNA

in J. tasmanica were checked using custom designed


3

3
3

3
primers and a touchdown PCR with an annealing
tmRNA

temperature of 50°C for nine cycles followed by a de-


crease in annealing temperature to 45°C for 25 cycles
1

1
1

(Table S2 in the Supporting Information).


tRNA

For Sanger sequencing, the cox1 and rbcL genes


29

29

29

29
29

29

were sequenced for various host and parasite combi-


nations. These samples were used to confirm para-
2 (moeB, ycf46)

2 (moeB, ycf46)
Pseudogenes

site and host samples from the HTS samples and as


a comparison to available data from other sequenced
Abbreviations: ncRNA, noncoding RNA; rRNA, ribosomal RNA; tmRNA, transfer-­messenger RNA; tRNA, transfer RNA.

Janczewskia species. The GazF1 and GazR1 primers


were used to amplify partial cox1 (Saunders, 2005),
(Ψ)

0
0

and various primer combinations (F2-­R1452, F7-­RrbcS


Start, F57-­rbclrevNEW, and F765-­R1442) were used to
Unclassified

amplify rbcL (Díaz-­Tapia et al., 2018; Freshwater & Rue-


ness, 1994; Gavio & Fredericq, 2002; Kim et al., 2010;
ORFs

Saunders & Moore, 2013; Wang et al., 2000). The an-


20

20

26
21
19

17

nealing temperature for all PCRs was 45°C, following


Preuss and Zuccarello (2018).
Protein coding

For the plastid genome phylogeny, a total of 193


genes (w/o

concatenated plastid protein sequences were extracted


ORFs)a

and aligned with all available closely related plastid ge-


194a

194a

195

194
194

195

nomes in the Rhodomelaceae (Table S3 in the Support-


ing Information) using MAFFT (Katoh & Standley, 2013)
content (%)

in Geneious. Vertebrata spp. were used as outgroups


following Díaz-­Tapia et al. (2019).
70.2

70.2
70.8

70.8

70.9
70.8

For the cox1 and rbcL alignment, over 300 sequences


A-­T

within Laurencia or the Laurencieae were downloaded


from GenBank. All unique sequences were selected
Plastome
size (bp)

173,808

172,933

176,465

173,028

based on maximum likelihood (ML) phylogenies in IQ-­


171,833
173,810

TREE (Nguyen et al., 2015), and the longest unique se-


quence per taxon was kept. Chondria spp. were used
as an outgroup for the single gene alignments based
Janczewskia tasmanica

Laurencia catarinensis

on Díaz-­Tapia et al. (2017). Samples sequenced using


(on L. catarinensis)
Janczewskia tasmanica

both HTS and Sanger sequencing produced identical


Including pseudogenes.
Laurencia obtusa
verruciformis
(on L. sp. A25)

sequences for each sample, and only one sequence


Laurencia elata
(on L. elata)

Janczewskia

was included in the phylogenetic analyses.


Suitable models of amino acid evolution were se-
TA B L E 1

Parasites

Non-­host

lected based on the Bayesian information criterion


Hosts

(BIC) using ModelFinder in IQ-­TREE (Kalyaanamoorthy


et al., 2017). The model JTTCDMut+F+I+G4 was used
a
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 953

for the multi-­gene phylogeny. The model TIM3+F+G4 coding genes, including all photosynthetic genes (Ta-
was used for the first, HKY+F+I for the second, and bles 1 and 2; Table S3). The plastid genomes of the two
TN+F+I+G4 for the third codon position in cox1. The samples of J. tasmanica were identical with a slight dif-
model TIM2+F+I+G4 was used for first and second ference in their complete size (173,808 vs. 173,810 bp).
codon position, and K3Pu+F+I+G4 was used for the This is the very first time a gene-­rich plastid genome of
third codon position, in rbcL. Maximum likelihood to- a parasitic red alga has been documented.
pologies were constructed, and branch support was Comparison between the six complete plastid ge-
determined using 1000 replicates for non-­parametric nomes of Janczewskia spp. and Laurencia spp. show a
bootstrap (Felsenstein, 1985), Ultrafast bootstrap conserved architectural structure with many similarities
(Hoang et al., 2018), and Shimodaira–­ Hasegawa-­ including genome size (171,833–­176,465 bp), number
approximate likelihood-­ ratio test (SH-­aLRT; Guindon of genes (194–­195), AT bias, and gene order (Table 1,
et al., 2010) analyses. Figure S1 in the Supporting Information). The plastid
Monophyly and multiple independent origins of the genomes are so similar that 90% of all shared protein
parasitic species were tested using the approximately coding genes (194) are the same length (Table S5 in
unbiased test (AU test) in IQ-­ TREE. An alternative the Supporting Information). Despite the great simi-
topology was created with three groups: parasites, larities between the plastid genomes of parasites and
free-­living red algae, and outgroup in Mesquite v3.70 free-­living algae, the protein coding genes ycf46 (un-
(Maddison & Maddison, 2021). characterized ATPase family AAA domain-­containing
Key parameter settings (e.g., random starting seed protein) and moeB (Molybdopterin biosynthesis protein)
number) of phylogenetic analyses to increase data re- are pseudogenes in J. tasmanica but not in J. verruci-
producibility can be found in Table S4 in the Supporting formis. In J. tasmanica, there is a frameshift mutation in
Information. the moeB gene caused by an indel and a stop codon
Synteny of the plastid genomes was first checked produced by a transversion, and a frameshift caused
using a progressive Mauve alignment (Darling by an indel are present in ycf46 (Table 1). In addition,
et al., 2004), and the plastid genomes were visualized the protein coding gene asnB (asparagine synthetase
linearly in OGDRAW (Greiner et al., 2019). B) is present in J. verruciformis (and L. obtusa) but not
For comparative analysis, plastid genomes were in J. tasmanica or any other compared Laurencia spp.
ordered by genome size/gene content as functional (Table S5).
and physical genome reduction should reflect different
transitional stages. Unannotated tRNAs were added to
the plastid genome of Choreocolax polysiphoniae and Tentative patterns of plastid genome
missing genes were manually checked to see if they evolution in parasitic red algae
were present in C. polysiphoniae and Harveyella mira-
bilis. We reannotated any missing protein coding genes Plastid genomes were compared in order of decreas-
by identifying their potential position and reading frame ing size and gene content using the two Janczews-
from closely related species. The new reading frame kia parasite species and previous data: Choreocolax
was translated to amino acids and added to the anno- polysiphoniae (Salomaki et al., 2015), Harveyella mi-
tation if BLAST hits were found. rabilis (Salomaki & Lane, 2019), and Pterocladiophila
hemisphaerica (Preuss et al., 2020). Two new anno-
tated hypothetical proteins were identified (ycf38 and
RESU LT S ycf19), and four non-­ annotated genes (ycf21, rpl32,
rpl33, and rpl35) were added to the plastid genome
Plastid genome characteristics and of C. polysiphoniae. In contrast, three genes (dnaB,
comparisons fabH, and rpl35) were missing in H. mirabilis and not
in the smaller plastid genomes of C. polysiphoniae and
Six complete and one draft plastid genome of three P. hemisphaerica. All compared genes in the reduced
parasite and host combinations and one close free-­ plastids had missing photosystem I and II genes, loss of
living relative were assembled. Two of the complete all the cytochrome b/f complex genes except for petF,
parasite plastid genomes are from Janczewskia tas- and loss of ATP synthase genes and Calvin–­Benson–­
manica, which we observed growing on two different Bassham cycle genes. There were conservation and
host species: Laurencia elata and an unidentified Lau- reduction in number of rpo, rps, and rpl genes in the
rencia sp. (here called A25; Table 1). The mean cover- parasites with the most reduced plastids (Choreoco-
age of the assembled plastid genomes varied between lax and Pterocladiophila; Table 2). The plastid genome
428 and 2585× and generated between 501,479 and of J. tasmanica indicates the first potentially reductive
2,949,102 reads. The plastid genomes of the parasites changes, as it contains two pseudogenes (moeB and
J. tasmanica and J. verruciformis are large, 172,933 ycf46) that were intact in J. verruciformis yet missing in
and 173,810 bp, and gene rich, with 194 and 195 protein the other parasites (Table 2). Major changes occurred
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
954 |    PREUSS et al.

TA B L E 2 Comparison of protein coding genes in parasitic red algae ordered from largest plastid genome (i.e., largest number of genes)
to smallest (i.e., with fewest number of genes; Table 3).

Janczewskia Janczewskia Harveyella Choreocolax Pterocladiophila


verruciformis tasmanica mirabilis polysiphoniae hemisphaerica
CDS
Photosystem I
psa genes (11)
Photosystem II
psb genes (19)
Cytochrome b/f complex
pet genes (10) 1 (petF) 1 (petF) 1 (petF)
ATP synthase
atp genes (8)
RuBisCO
rbcL gene (1)
RNA polymerase
rpo genes (5) 4 4
Ribosomal proteins—­SSU
rps genes (19) 18 16 16
Ribosomal proteins— ­LSU
rpl genes (28) 26a 24 21
Other conserved ORFs
ycf genes (26) Ψ ycf45 6 3 3
Other genes (67)c Ψ moeB 28a (Ψ gltBb) 27 (Ψ gltBb) 23
Note: Genes are grouped based on functional groups shown in OGDRAW, and total number of genes are seen in brackets. Dark gray cells indicate all genes
are present and light gray cells indicate gene loss with the total number of remaining genes given. Blank region indicates no gene for that functional group was
found in the plastid genome. Pseudogenes are indicated by Ψ.
a
dnaB, fabH and rpl35 are lost in H. mirabilis but not in C. polysiphoniae and P. hemisphaerica.
b
Following Salomaki and Lane (2019).
c
asnB was excluded from the comparison as it is present in J. verruciformis and absent in J. tasmanica and some free-­living red algae.

in H. mirabilis: the plastid genome dramatically reduced to the phylogenies, that these species had evolved at
in size due to high gene loss, but synteny was main- least twice within Laurencia (Figure 1). The plastid ge-
tained in the remaining genes compared to free-­living nomes of the two samples of J. tasmanica were dis-
red algae. Both C. polysiphoniae and P. hemisphaerica tantly related to their two different host species, L. elata
showed structural rearrangements, along with gene and Laurencia sp. A25. Janczewskia verruciformis was
loss, in their plastid genomes, which were not observed closely related to the non-­host species L. obtusa and
in Janczewskia and Harveyella plastid genomes. distantly related to its host L. catarinensis (Figure 1).
The 663 bp alignment of the cox1 gene contained
61 taxa including four different Janczewskia parasites
Phylogeny confirms multiple and their associated host species. The ML phylogeny
parasite origins in Laurencia did not support the monophyly of Janczewskia, as the
parasites evolved independently multiple times within
The 46,454 amino acid alignment of 193 genes in- Laurencia (Figure 2). Most Janczewskia species were
cluded 21 taxa, including Janczewskia tasmanica and distantly related to their hosts, except for J. morimo-
J. verruciformis and their Laurencia host species. The toi which was closely related to its host L. nipponica.
ML phylogeny showed four main lineages within Lau- The Australian parasite J. tasmanica was distantly
rencieae that corresponded to two recognized genera related to its two host species, as shown earlier,
(Laurenciella and Palisada), an unnamed lineage (Lau- and formed an unsupported grouping with two non-­
rencieae sp.), and a lineage including both Laurencia hosts (L. karachiana MK796229 and Laurencia sp.
and Janczewskia. All nodes within the phylogeny were KX258826). The European parasite J. verruciformis
fully supported except for one (L. catarinensis and J. tas- was nested in a group with specimens identified as
manica). Alternative topologies rejected (p = 0.000) the L. obtusa (our data and KX258828) from France and
monophyly of Janczewskia and confirmed, in addition L. viridis (KF492757) from the Canary Islands, Spain,
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 955

supported by non-­parametric and ultrafast bootstrap, was observed (two pseudogenes), and despite their
and distantly related to its host L. catarinensis. The similarity, these two Janczewskia species evolved in-
Hawaiian parasite J. hawaiiana grouped with boot- dependently. Comparing all circular-­mapping parasite
strap support to an unidentified non-­host Laurencia plastid genomes shows some tentative evolutionary
sp. (GU223893) and not its host species, L. mcder- changes that are indicative of an order of change in
midiae (Figure 2). these genomes during the transition from free-­living to
The 1467 bp alignment of the rbcL gene contained parasitic. Also, the studied Janczewskia species do not
84 taxa including three Janczewskia parasites and support its generic status, as the parasites evolved in-
their host species, when available. A third host species dependently within Laurencia.
for J. tasmanica, Laurencia elata, was included, too. Janczewskia was established based on the mor-
The ML phylogeny supported that the three included phological characters of J. verruciformis (Solms-­
Janczewskia species evolved independently within Laubach, 1877). Molecular data do not confirm the
Laurencia (Figure 3). Janczewskia tasmanica formed monophyly of the genus and show that all four inves-
a strongly supported grouping with the non-­host spe- tigated species (J. hawaiiana, J. morimotoi, J. tasman-
cies L. karachiana (MK796228) from Pakistan. The ica, and J. verruciformis) evolved independently within
parasite was distantly related to L. elata, Laurencia Laurencia (Kurihara et al., 2010; this study). Placing
sp. A25, and L. tasmanica, which were closely related parasitic red algae into the genus of its closest rela-
to each other. The European parasite J. verruciformis tive to reflect their evolutionary origin instead of para-
was nested within a clade of L. obtusa sequences sitic genera has become a common practice (Freese
with strong phylogenetic support and was distantly & Lane, 2021; Gurgel et al., 2018; Preuss et al., 2017;
related to its host L. catarinensis. The Japanese par- Preuss & Zuccarello, 2018; Preuss & Zuccarello, 2019;
asite J. morimotoi formed a strongly supported group Silva, 1991). While not all Janczewskia species have
with two specimens in GenBank submitted as “Chon- been sequenced, we believe, based on available data
drophycus spp.” (DQ787585, MH388594) from South and to preserve the monophyletic status of Laurencia
Korea and Japan, respectively (Figure 3). No rbcL that now contains several polyphyletic Janczewskia
sequences from its host L. nipponica were available, spp., that the generic name for these parasites should
and therefore it is uncertain if these topologies are be changed. Also, transferring the generitype J. verru-
congruent between mitochondrial and plastid data ciformis, based on our data, would leave the yet un-
sets. sequenced species as generic orphans. The slight
possibility that a minority of these unsequenced orphan
parasite species growing on other genera (Acantho-
D I SCUS S I O N phora, Chondria, and Osmundea; Preuss et al., 2017)
might be transferred from Laurencia in the future is
Evolution and taxonomic placement of of course a possibility, but it is clear that the genus
Janczewskia Janczewskia cannot be maintained. During this writ-
ing, morphological work proposed the transfer of three
Our study demonstrated, for the first time, that parasitic Janczewskia species (J. gardneri, J. moriformis, and
red algae can contain a gene-­rich plastid genome, as J. lappacea) into a new parasitic genus, Heterojancze-
seen in Janczewskia verruciformis, which is compa- wskia (Nam, 2022). We believe these changes ignored
rable to the standard plastid genome in free-­living red previous studies showing that Janczewskia was not
algae. Despite the gene richness of the plastid genome monophyletic (Kurihara et al., 2010) and was nested
of J. tasmanica, evidence of early stages of degradation in Laurencia. The shared morphological characters of

TA B L E 3 Comparison of plastid genome features between five parasitic red algal species ordered from largest plastid genome with
highest number of genes to smallest.

Janczewskia Janczewskia Harveyella Choreocolax Pterocladiophila


verruciformis tasmanica mirabilis polysiphoniae hemisphaerica
Size (bp) 172,933 173,808–­173,810 90,654 90,243 68,701
A-­T content (%) 70.2 70.8 76.5 79.5 74.2
CDS 195 194 84 75 70
Pseudogenes 0 2 1 1 0
tRNAs 29 29 24 25 26
Rearrangements no no no yes yes
Note: Gene order of free-­living red algae was compared to parasites and indicated if any rearrangements were present. Features included total plastid genome
size, A-­T content, protein coding genes (CDS) without unclassified ORFs, pseudogenes, transfer RNAs (tRNAs), and plastid genome rearrangements. A-­T
content increases with loss of CDSs.
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
956 |    PREUSS et al.

the parasites with their host (e.g., J. gardneri) indicates Laurencia morimotoi (Tokida) M.Preuss, Diaz-­
the share evolutionary origin with their host and are Tapia, Verbruggen & Zuccarello comb. nov.
not a separate genus as is often the case for red algal Basionym: Janczewskia morimotoi Tokida, J. Jap.
parasites. Bot. 21: 127, figs. 1–­6 (Tokida, 1947).
We propose that all species in the genus Janczews- Heterotypic synonym: Janczewskia tokidae
kia and Heterojanczewskia be transferred to the genus Saito (1971).
Laurencia: Laurencia parasitica M.Preuss, Diaz-­ Tapia, Ver-
Laurencia australis M.Preuss, Diaz-­Tapia, Verbrug- bruggen & Zuccarello nom. nov.
gen & Zuccarello nom. nov. Replaced name: Janczewskia moriformis Setch.,
Replaced name: Janczewskia tasmanica Falkenb., Univ. Calif. Publ. Bot. 6: 11, 12, 21, pl. 1, figs. 1-­3; pl. 3,
in Schmitz and Falkenberg (1897), Die natürlichen figs. 20, 21; pl. 4, figs. 22, 23; pl. 5, fig. 24 (1914).
Pflanzenfamilien nebst ihren Gattungen und wichtig- Homotypic synonym: Heterojansczewskia mori-
eren Arten insbesondere den Nutzpflanzen unter Mit- formis (Setch.) K.W.Nam. Korean J. Environ. Biol 40:
wirkung zahlreicher hervorragender Fachgelehrten, 305 (2022).
Teil 1, Abteilung 2, p. 432, fig. 243 C. Etymology: (Latin = parasitic) refers to the parasitic
Heterotypic synonym: Janczewskia australis life mode of this alga.
Falkenb. ex Reinbold (1899) nom. ined. Note: The existence of Laurencia moriformis
Etymology: (Latin = southern) refers to the country ­Kützing (1865) precludes the transfer of Janczewskia
it is found in. moriformis Setch. into Laurencia, which would result in
Note: The existence of Laurencia tasmanica Hook.f. a later homonym.
& Harv., in Harvey (1849) precludes the transfer of Laurencia ramiformis (C.F.Chang & B.M.Xia)
Janczewskia tasmanica Falkenb. into Laurencia, which M.Preuss, Diaz-­Tapia, Verbruggen & Zuccarello comb.
would result in a later homonym. nov.
Laurencia hawaiiana (Apt) M.Preuss, Diaz-­Tapia, Basionym: Janczewskia ramiformis C.F.Chang &
Verbruggen & Zuccarello comb. nov. B.M.Xia, Stud. Mar. Sin. 14: 120, 126, fig. 1: 1–­8, fig. 2:
Basionym: Janczewskia hawaiiana Apt, Phycologia 1–­6 (Chang & Xia, 1978).
26: 328, 329, figs. 1–­7 (Apt, 1987). Laurencia solmsii (Setch. & Guernsey) M.Preuss,
Laurencia janczewskii M.Preuss, Diaz-­Tapia, Ver- Diaz-­Tapia, Verbruggen & Zuccarello comb. nov.
bruggen & Zuccarello nom. nov. Basionym: Janczewskia solmsii Setch. & Guernsey,
Replaced name: Janczewskia gardneri Setch. in Setchell, Univ. Calif. Publ. Bot. 6: 9, 10, 11, 20, pl.
& Guernsey, in Setchell Univ. Calif. Publ. Bot. 6: 12, 2, figs. 7, 8, pl. 3, figs. 17–­19; pl. 5, figs. 26, 27 (1914).
13, 14, 21, pl. 1, figs. 4–­6; pl. 3, figs. 15, 16; pl. 5, fig. Laurencia teysmannii (Weber Bosse) M.Preuss,
25 (1914). Diaz-­Tapia, Verbruggen & Zuccarello comb. nov.
Homotypic synonym: Heterojansczewskia gardneri Basionym: Janczewskia teysmannii Weber Bosse,
(Setch. & Guernsey) K.W.Nam. Korean J. Environ. Biol Siboga-­Expeditie Monographie p. 348, fig. 133 (Weber-­
40: 305 (2022). van Bosse, 1923).
Etymology: “janczewskii” of Janczewski, in refer- Laurencia verruciformis (Solms-­Laubach) M.Pre-
ence to the previous designated genus name Jancze- uss, Diaz-­Tapia, Verbruggen & Zuccarello comb. nov.
wskia Solms-­Laubach (1877). Basionym: Janczewskia verruciformis Solms-­
Note: The existence of Laurencia gardneri Hollenb., Laubach, Mém. Soc. Natl. Sci. Nat. Math. Cherbg. 21:
in Smith & Hollenberg (1943) precludes the transfer of 207, 210, pl. 3: figs. 1–­18 (1877).
Janczewskia gardneri Setch. & Guernsey into Lauren-
cia, which would result in a later homonym.
Laurencia lappacea (Setch.) M.Preuss, Diaz-­Tapia, Plastid genome degradation and evolution
Verbruggen & Zuccarello comb. nov.
Basionym: Janczewskia lappacea Setch., in Setch- The discovery of gene-­rich, and barely reduced, plas-
ell, Univ. Calif. Publ. Bot. 6: 14–­16, 21, 22; pl. 6, figs. tid genomes in red algal parasites is important as it
28–­32 (Setchell, 1914). indicates that fine-­scale patterns of plastid genome
Homotypic synonym: Heterojansczewskia lappa- degradation could be seen in parasites of different evo-
cea (Setch.) K.W.Nam. Korean J. Environ. Biol 40: 305 lutionary ages or dependencies on their host species.
(2022). The first evolutionary change that can be observed, in
Laurencia meridionalis (M.T.Martin & Pocock) Laurencia australis (as Janczewskia tasmanica), is a
M.Preuss, Diaz-­Tapia, Verbruggen & Zuccarello comb. pseudogenization of two genes (moeB and ycf46). This
nov. pseudogenization was not observed in L. verruciformis
Basionym: Janczewskia meridionalis M.T.Martin & (as J. verruciformis). In flowering parasitic plants, pre-
Pocock, J. Linn. Soc. London, Bot. 55: 62, figs. 3a, b dominately ndh photosynthetic genes initially became
(Martin & Pocock, 1953). pseudogenes and were subsequently lost (Wicke
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 957

F I G U R E 1 Maximum likelihood topology of 193 concatenated plastid gene amino acid sequences from the two parasites Janczewskia
tasmanica and J. verruciformis and closely related species within the Rhodomelaceae. Host and parasite combinations are indicated by
arrows. Outgroups of Vertebrata spp. were removed to facilitate presentation. Scale bar represents substitution per site. Branch support
values are given as non-­parametric bootstrap/UFbootstrap/SH-­aLRT. Asterisk (*) represents 100% support.

F I G U R E 2 Maximum likelihood topology of the cox1 sequences of four Janczewskia species and their associated hosts. Newly
sequenced taxa are highlighted in bold, and arrows connect each parasite to its host species. Chondria spp. outgroups were removed to
facilitate presentation. Scale bar represents substitution per site. Branch support values are given as bootstrap/UFbootstrap/SH-­aLRT.
Values <85% bootstrap, <95% UFbootstrap, and < 80% SH-­aLRT test are not shown. Asterisk (*) represents 100% support.

et al., 2013; Yu et al., 2017). This gene family is not pre- The first two discovered pseudogenes, moeB and
sent in other red algal plastids (Lee et al., 2016; Preuss ycf46 in Laurencia australis, have been completely lost
et al., 2020, 2021). in the parasites Choreocolax polysiphoniae, Harveyella
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
958 |    PREUSS et al.

F I G U R E 3 Maximum likelihood topology of the rbcL gene of three Janczewskia parasites and some of their associated hosts and
related species. New sequenced taxa are highlighted in bold, and arrows connect each parasite to its host species. The outgroup of
Chondria spp. were removed to facilitate presentation. Branch support values are given as bootstrap/UFbootstrap/SH-­aLRT. Values <85%
bootstrap, <95% UFbootstrap, and <80% SH-­aLRT test are not shown. Asterisk (*) represents 100% support.

mirabilis, and Pterocladiophila hemisphaerica (Pre- that the loss of ycf46 is not linked to parasitism but an
uss et al., 2020; Salomaki et al., 2015; Salomaki & easily lost function. More gene-­rich plastids and plastids
Lane, 2019). The gene moeB produces an enzyme that with different levels of degradation from various parasitic
catalyzes ATP in the molybdopterin biosynthetic path- algae are needed to determine if there are any regular
way (Leimkühler et al., 2001), which is not connected to patterns in gene pseudogenization and loss.
photosynthetic capability, and this function might not be Multiple models for the order of gene loss in para-
essential in parasites. In contrast, ycf46 is predicted to sitic plants have been proposed, which overall show
play a role in the uptake and utilization of CO2 during pho- more similarities than differences (e.g., Barrett &
tosynthesis (Jiang et al., 2015). The loss of ycf46 in the Davis, 2012; Naumann et al., 2016: Wicke et al., 2016).
plastid genome is not unusual in free-­living red algae and For example, a five-­stage model of gene loss has been
has been observed in the freshwater orders Balbiania- proposed, where ndh genes are lost first, followed
les, Batrachospermales, and Thoreales (Cho et al., 2018; by photosynthetic genes (psa, psb, and pet), then
Evans et al., 2019; Paiano et al., 2018). This could indicate PEP genes (rpo), then ATP synthase (atp), and lastly
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 959

other functions (Barrett & Davis, 2012). The model their host to complete their life cycle. This hemipa-
of Barrett and Davis was then modified by allowing rasitic life strategy may be a common characteristic in
for these five stages to overlap (Graham et al., 2017). the early stages of many red algal parasites that then
Keeping in mind that gene loss in not linear (Graham evolve increased host dependency and subsequent in-
et al., 2017), one gene per million years is estimated creased plastid genome reduction.
to be lost in parasitic plants (broomrapes; Cusimano This is the first observation of not only gene-­rich
& Wicke, 2016). Our plastid gene loss comparison plastid genomes of red algal parasites but also the early
of parasitic red algae, while still preliminary, shows stages of plastid genome degeneration. Future studies
gene categories associated with photosynthesis are of other parasitic red algae may show patterns that are
missing (except for petF), while reduced genomes useful for understanding the changes that occur in the
have less but still contain gene categories associated transition between free-­living and parasitic life strate-
with non-­ photosynthetic functions (e.g., ribosomal gies in these fascinating organisms.
proteins: rps, rpl; Table 2). So loss of autotrophic abil-
ity may be the initial change that is seen in the tran- AU T H O R C O N T R I B U T I O N S
sition from free-­living to parasitism. Similar orders of Maren Preuss: Conceptualization (lead); data cura-
gene loss have been observed in plastid genomes of tion (lead); formal analysis (lead); funding acquisition
autotrophic lineages of apicomplexans (e.g., Plasmo- (lead); investigation (lead); methodology (lead); project
dium), chrysophytes, and diatoms (Kim et al., 2020). administration (lead); resources (equal); software (lead);
The minimal plastid size, before complete loss of the validation (lead); visualization (lead); writing –­original
plastid genome, has been postulated in the highly re- draft (lead); writing –­review and editing (lead). Pilar
duced plastid genome of the chrysophyte “Spumella” Díaz-­Tapia: Data curation (supporting); funding acqui-
(Dorrell et al., 2019). Loss of all photosynthetic genes sition (supporting); resources (equal); writing –­review
in independently evolved parasites are an indication and editing (supporting). Heroen Verbruggen: Fund-
that these parasitic systems can be used to identify ing acquisition (supporting); resources (equal); software
patterns of plastid degradation that are applicable to (supporting); writing –­review and editing (supporting).
other parasites evolved from photosynthetic lineages. Giuseppe C. Zuccarello: Funding acquisition (support-
ing); project administration (supporting); writing –­origi-
nal draft (supporting); writing –­review and editing (lead).
Laurencia australis and L.
verruciformis are obligate hemiparasites AC K N O​W L E​D G E​M E N T S
We thank Michael Wynne and Craig Schneider for
Obligate parasites cannot live independently from their advice on nomenclature. This work was funded by a
host and can be further divided into hemiparasites, Marsden Fast Start (Grant no 19-­ VUW-­006) by the
which have retained their photosynthetic capability, Royal Society Te Apārangi, the Australian Biological
and holoparasites, which have lost photosynthesis Resources Study (Activity 4-­G046WSD) and Xunta de
completely. Our study showed the pigmented para- Galicia (grants ED481D/ 2017/011 and 03IN858A2019-
sites Laurencia australis and L. verruciformis retained ­1630129). Phylogenetic analyses were performed on
all photosynthetic genes and that these parasites are the Rāpoi High-­Performance Computing Facility of Vic-
potentially capable of carrying out photosynthesis inde- toria University of Wellington, New Zealand.
pendently. Early culture work showed that Janczewskia
gardneri incorporated NaH14CO3 when attached and ORCID
unattached to its host Laurencia spectabilis (now Os- Maren Preuss https://orcid.
mundea spectabilis), which led to the assumptions that org/0000-0002-8147-5643
parasite plastids were photosynthetic (Court, 1980). Pilar Díaz-­Tapia https://orcid.
Later our understanding of parasite plastid origin org/0000-0003-4680-4867
changed, when it was shown that parasites transferred Heroen Verbruggen https://orcid.
material, including plastids, into host cells via sec- org/0000-0002-6305-4749
ondary pit connections, and it was assumed that the Giuseppe C. Zuccarello https://orcid.
host plastids were retained in the parasite cells (Goff org/0000-0003-0028-7227
& Coleman, 1995) and that parasite plastids were lost
or non-­functional. Despite the photosynthetic capabil- REFERENCES
ity of the parasite, these organisms rely on the host to Apt, K. E. (1987). A new species of Janczewskia (Rhodomelaceae,
complete their development as shown in J. morimotoi Rhodophyta) from the Hawaiian islands. Phycologia, 26(3),
(Nonomura, 1979). The presence of all photosynthetic 328–­333.
Banerjee, A., Schneider, A. C., & Stefanović, S. (2022). Plastid ge-
genes in the plastid genomes and pigmentation in the nomes of the hemiparasitic genus Krameria (Zygophyllales)
two parasites L. australis and L. verruciformis suggests are intact and exhibit little relaxation in selection. International
that these are obligate hemiparasites but depend on Journal of Plant Sciences, 183(5), 393– ­403.
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
960 |    PREUSS et al.

Barrett, C. F., & Davis, J. I. (2012). The plastid genome of the myco- Goff, L. J., Moon, D. A., Nyvall, P., Stache, B., Mangin, K., &
heterotrophic Corallorhiza striata (Orchidaceae) is in the rela- Zuccarello, G. (1996). The evolution of parasitism in the red
tively early stages of degradation. American Journal of Botany, algae: Molecular comparisons of adelphoparasites and their
99(9), 1513–­1523. hosts. Journal of Phycology, 32(2), 297–­312.
Beck, N., & Lang, B. (2010). MFannot, organelle genome annotation Graham, S. W., Lam, V. K., & Merckx, V. S. (2017). Plastomes on the
websever. Université de Montréal QC. edge: The evolutionary breakdown of mycoheterotroph plastid
Bolger, A. M., Lohse, M., & Usadel, B. (2014). Trimmomatic: A flexi- genomes. New Phytologist, 214(1), 48–­55.
ble trimmer for Illumina sequence data. Bioinformatics, 30(15), Greiner, S., Lehwark, P., & Bock, R. (2019). OrganellarGenomeDRAW
2114–­2120. (OGDRAW) version 1.3.1: Expanded toolkit for the graphical
Chang, C. F., & Xia, B. M. (1978). Studies on the parasitic red algae visualization of organellar genomes. Nucleic Acids Research,
of China. Studia Marina Sinica, 14, 119–­127. 47(W1), W59–­W64.
Cho, C. H., Choi, J. W., Lam, D. W., Kim, K. M., & Yoon, H. S. (2018). Guindon, S., Dufayard, J. F., Lefort, V., Anisimova, M., Hordijk, W.,
Plastid genome analysis of three Nemaliophycidae red algal & Gascuel, O. (2010). New algorithms and methods to estimate
species suggests environmental adaptation for iron limited maximum-­likelihood phylogenies: Assessing the performance
habitats. PLoS ONE, 13(5), e0196995. of PhyML 3.0. Systematic Biology, 59(3), 307–­321.
Court, G. J. (1980). Photosynthesis and translocation studies of Guiry, M. D., & Guiry, G. M. (2022). AlgaeBase. World-­Wide
Laurencia spectabilis and its symbiont Janczewskia gardneri Electronic Publication, National University of Ireland.
(Rhodophyceae). Journal of Phycology, 16(2), 270–­279. Gurgel, C. F. D., Norris, J. N., Schmidt, W. E., Le, H. N., & Fredericq,
Cusimano, N., & Wicke, S. (2016). Massive intracellular gene transfer S. (2018). Systematics of the Gracilariales (Rhodophyta) in-
during plastid genome reduction in nongreen Orobanchaceae. cluding new subfamilies, tribes, subgenera, and two new gen-
New Phytologist, 210(2), 680– ­693. era, Agarophyton gen. nov. and Crassa gen. nov. Phytotaxa,
Darling, A. C., Mau, B., Blattner, F. R., & Perna, N. T. (2004). Mauve: 374(1), 1–­23.
Multiple alignment of conserved genomic sequence with rear- Harvey, W. H. (1849). Nereis australis, or algae of the southern
rangements. Genome Research, 14(7), 1394–­1403. ocean: Being figures and descriptions of marine plants, col-
de Koning, A. P., & Keeling, P. J. (2006). The complete plastid ge- lected on the shores of the Cape of Good Hope, the extra-­
nome sequence of the parasitic green alga Helicosporidium tropical Australian colonies, Tasmania, New Zealand, and the
sp. is highly reduced and structured. BMC Biology, 4(1), 12. Antarctic regions; deposited in the herbarium of the Dublin
Díaz-­Tapia, P., Maggs, C. A., Macaya, E. C., & Verbruggen, H. university. Reeve.
(2018). Widely distributed red algae often represent hidden Hoang, D. T., Chernomor, O., Von Haeseler, A., Minh, B. Q., &
introductions, complexes of cryptic species or species with Vinh, L. S. (2018). UFBoot2: Improving the ultrafast boot-
strong phylogeographic structure. Journal of Phycology, 54(6), strap approximation. Molecular Biology and Evolution, 35(2),
829–­839. 518–­522.
Díaz-­Tapia, P., Maggs, C. A., West, J. A., & Verbruggen, H. (2017). Jiang, H. B., Song, W. Y., Cheng, H. M., & Qiu, B. S. (2015). The
Analysis of chloroplast genomes and a supermatrix inform re- hypothetical protein ycf46 is involved in regulation of CO2 uti-
classification of the Rhodomelaceae (Rhodophyta). Journal of lization in the cyanobacterium Synechocystis sp. PCC 6803.
Phycology, 53(5), 920–­937. Planta, 241, 145–­155.
Díaz-­Tapia, P., Pasella, M. M., Verbruggen, H., & Maggs, C. A. Kalyaanamoorthy, S., Minh, B. Q., Wong, T. K., Von Haeseler, A.,
(2019). Morphological evolution and classification of the red & Jermiin, L. S. (2017). ModelFinder: Fast model selection
algal order Ceramiales inferred using plastid phylogenomics. for accurate phylogenetic estimates. Nature Methods, 14(6),
Molecular Phylogenetics and Evolution, 137, 76–­85. 587– ­589.
Dorrell, R. G., Azuma, T., Nomura, M., Audren de Kerdrel, G., Paoli, Katoh, K., & Standley, D. M. (2013). MAFFT multiple sequence
L., Yang, S., & Kamikawa, R. (2019). Principles of plastid reduc- alignment software version 7: Improvements in perfor-
tive evolution illuminated by nonphotosynthetic chrysophytes. mance and usability. Molecular Biology and Evolution, 30(4),
Proceedings of the National Academy of Sciences, 116(14), 772–­780.
6914– ­6923. Kayama, M., Maciszewski, K., Yabuki, A., Miyashita, H.,
Evans, J. R., St. Amour, N., Verbruggen, H., Salomaki, E. D., & Vis, M. Karnkowska, A., & Kamikawa, R. (2020). Highly reduced
L. (2019). Chloroplast and mitochondrial genomes of Balbiania plastid genomes of the non-­ photosynthetic dictyochophy-
investiens (Balbianiales, Nemaliophycidae). Phycologia, 58(3), ceans Pteridomonas spp.(Ochrophyta, SAR) are retained for
310–­318. tRNA-­Glu-­based organellar heme biosynthesis. Frontiers in
Felsenstein, J. (1985). Confidence limits on phylogenies: An ap- Plant Science, 11, 602455.
proach using the bootstrap. Evolution, 39(4), 783–­791. Kim, J. I., Jeong, M., Archibald, J. M., & Shin, W. (2020). Comparative
Freese, J. M., & Lane, C. E. (2021). Reorganizing parasitic plastid genomics of non-­photosynthetic chrysophytes: Genome
Delesseriaceae: Taxonomic revision of Asterocolax. Phytotaxa, reduction and compaction. Frontiers in Plant Science, 11,
525(2), 124–­136. 572703.
Freshwater, D. W., & Rueness, J. (1994). Phylogenetic relationships Kim, S. Y., Weinberger, F., & Boo, S. M. (2010). Genetic data hint at a
of some European Gelidium (Gelidiales, Rhodophyta) species, common donor region for invasive Atlantic and pacific popula-
based on rbcL nucleotide sequence analysis. Phycologia, tions of Gracilaria vermiculophylla (Gracilariales, Rhodophyta).
33(3), 187–­194. Journal of Phycology, 46(6), 1346–­1349.
Gavio, B., & Fredericq, S. (2002). Grateloupia turuturu Kurihara, A., Abe, T., Tani, M., & Sherwood, A. R. (2010). Molecular
(Halymeniaceae, Rhodophyta) is the correct name of the non-­ phylogeny and evolution of red algal parasites: A case study of
native species in the Atlantic known as Grateloupia doryphora. Benzaitenia, Janczewskia, and Ululania (Ceramiales). Journal
European Journal of Phycology, 37(3), 349–­359. of Phycology, 46(3), 580–­590.
Goff, L. J., Ashen, J., & Moon, D. (1997). The evolution of para- Kützing, F. T. (1865). Tabulae phycologicae; oder, Abbildungen der
sites from their hosts: A case study in the parasitic red algae. Tange (Vol. XV). Gedruckt auf kosten des Verfassers.
Evolution, 51(4), 10681078. Laslett, D., & Canback, B. (2004). ARAGORN, a program to de-
Goff, L. J., & Coleman, A. W. (1995). Fate of parasite and host or- tect tRNA genes and tmRNA genes in nucleotide sequences.
ganelle DNA during cellular transformation of red algae by their Nucleic Acids Research, 32(1), 11–­16. https://doi.org/10.1093/
parasites. The Plant Cell, 7(11), 1899–­1911. nar/gkh152
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
PLASTID EVOLUTION IN JANCZEWSKIA     | 961

Lee, J., Cho, C. H., Park, S. I., Choi, J. W., Song, H. S., West, J. A., Prjibelski, A., Antipov, D., Meleshko, D., Lapidus, A., & Korobeynikov,
Bhattacharya, D., & Yoon, H. S. (2016). Parallel evolution of A. (2020). Using SPAdes de novo assembler. Current Protocols
highly conserved plastid genome architecture in red seaweeds in Bioinformatics, 70(1), e102.
and seed plants. BMC Biology, 14, 75. https://doi.org/10.1186/ Reinbold, T. (1899). Meeresalgen von Investigator Street (Süd-
s1291​5 -­016-­0299-­5 Australien), gesammelt von Miss Nellie Davey (Waltham,
Leimkühler, S., Wuebbens, M. M., & Rajagopalan, K. V. (2001). Honiton). Hedwigia, 38(1), 39–­51.
Characterization of Escherichia coli MoeB and its involvement Saito, Y. (1971). Two species of Janczewskia from Japan and their
in the activation of molybdopterin synthase for the biosynthesis systematic relationships. Proceedings of the International
of the molybdenum cofactor. Journal of Biological Chemistry, Seaweed Symposium, 7, 146–­149.
276(37), 34695–­34701. Salomaki, E. D., & Lane, C. E. (2019). Molecular phylogenetics sup-
Maddison, W. P., & Maddison, D. R. (2023). Mesquite: a modular ports a clade of red algal parasites retaining native plastids:
system for evolutionary analysis. Version 3.81. http://www. Taxonomy and terminology revised. Journal of Phycology,
mesqu​itepr​oject.org 55(2), 279–­288.
Martin, M. T., & Pocock, M. A. (1953). South African parasitic Salomaki, E. D., Nickles, K. R., & Lane, C. E. (2015). The ghost
Florideae and their hosts. 2. Some south African parasitic plastid of Choreocolax polysiphoniae. Journal of Phycology,
Florideae. Botanical Journal of the Linnean Society, 55(356), 51(2), 217–­221.
48–­64. Saunders, G. W. (2005). Applying DNA barcoding to red macroal-
Nam, K. W. (2022). Heterojanczewskia stat. nov. with an emenda- gae: A preliminary appraisal holds promise for future appli-
tion of generic delineation of Janczewskia (Rhodomelaceae, cations. Philosophical Transactions of the Royal Society, B:
Rhodophyta). Korean Journal of Environmental Biology, 40(3), Biological Sciences, 360(1462), 1879–­1888.
301–­306. Saunders, G. W., & Moore, T. E. (2013). Refinements for the ampli-
Naumann, J., Der, J. P., Wafula, E. K., Jones, S. S., Wagner, fication and sequencing of red algal DNA barcode and RedToL
S. T., Honaas, L. A., Ralph, P. E., Bolin, J. F., Maass, E., phylogenetic markers: A summary of current primers, profiles
Neinhuis, C., Wanke, S., & dePamphilis, C. W. (2016). and strategies. Algae, 28(1), 31–­43.
Detecting and haracterizing the highly divergent plastid Schmitz, F., & Falkenberg, P. (1897). Rhodomelaceae. In A. Engler
­g enome of the nonphotosynthetic parasitic plant Hydnora & K. Prantl (Eds.), Die natürlichen Pflanzenfamilien nebst ihren
visseri (Hydnoraceae). Genome Biology and Evolution, 8(2), Gattungen und wichtigeren Arten insbesondere den Nutzpflanzen
345–­3 63. unter Mitwirkung zahlreicher hervorragender Fachgelehrten, Teil
Ng, P. K., Lim, P. E., Kato, A., & Phang, S. M. (2014). Molecular 1, Abteilung 2. verlag von Wilhelm Engelmann.
evidence confirms the parasite Congracilaria babae Setchell, W. A. (1914). Parasitic Florideae (Vol. 6). University of
(Gracilariaceae, Rhodophyta) from Malaysia. Journal of California Press.
Applied Phycology, 26, 1287–­1300. Silva, P. C. (1991). Notes on the type specimens of red algal para-
Nguyen, L. T., Schmidt, H. A., Von Haeseler, A., & Minh, B. Q. (2015). sites described from California by WA Setchell. Taxon, 40(3),
IQ-­TREE: A fast and effective stochastic algorithm for estimat- 463–­470.
ing maximum-­likelihood phylogenies. Molecular Biology and Smith, G. M., & Hollenberg, G. J. (1943). On some Rhodophyceae
Evolution, 32(1), 268–­274. from the Monterey Peninsula. California. American Journal of
Nonomura, A. M. (1979). Development of Janczewskia morimotoi Botany, 30(3), 211–­223.
(Ceramiales) on its host Laurencia nipponica (Ceramiales, Solms-­Laubach, H. (1877). Note sur Janczewskia, nouvelle Floridée
Rhodophyceae). Journal of Phycology, 15(2), 154–­162. parasite du Chondria obtusa. Mémoires de la Société Des
Paiano, M. O., Del Cortona, A., Costa, J. F., Liu, S. L., Verbruggen, Sciences Naturelles et mathématiques de Cherbourg, 21,
H., De Clerck, O., & Necchi, O., Jr. (2018). Organization 209–­224.
of plastid genomes in the freshwater red algal order Tokida, J. (1947). Notes on some new or little known marine algae.
Batrachospermales (Rhodophyta). Journal of Phycology, Journal of Japanese Botany, 21, 127–­130.
54(1), 25–­3 3. Wang, H. W., Kawaguchi, S., Horiguchi, T., & Masuda, M. (2000).
Preuss, M., Nelson, W. A., & Zuccarello, G. C. (2017). Red algal Reinstatement of Grateloupia catenata (Rhodophyta,
parasites: A synopsis of described species, their hosts, dis- Halymeniaceae) on the basis of morphology and rbcL se-
tinguishing characters and areas for continued research. quences. Phycologia, 39(3), 228–­237.
Botanica Marina, 60(1), 13–­25. Weber-­ van Bosse, A. (1923). Liste des algues du Siboga, II:
Preuss, M., Verbruggen, H., West, J. A., & Zuccarello, G. C. (2021). Rhodophyceae, Seconde partie: Ceramiales. Siboga-­
Divergence times and plastid phylogenomics within the Expeditie Monographie, 59, 311–­392.
intron-­rich order Erythropeltales (Compsopogonophyceae, Wicke, S., Müller, K. F., de Pamphilis, C. W., Quandt, D., Wickett,
Rhodophyta). Journal of Phycology, 57(3), 1035–­1044. N. J., Zhang, Y., Renner, S. S., & Schneeweiss, G. M. (2013).
Preuss, M., Verbruggen, H., & Zuccarello, G. C. (2020). The Mechanisms of functional and physical genome reduction in
organelle genomes in the photosynthetic red algal para- photosynthetic and nonphotosynthetic parasitic plants of the
site Pterocladiophila hemisphaerica (Florideophyceae, broomrape family. The Plant Cell, 25(10), 3711– ­3725.
Rhodophyta) have elevated substitution rates and extreme Wicke, S., Müller, K. F., dePamphilis, C. W., Quandt, D., Bellot, S., &
gene loss in the plastid genome. Journal of Phycology, 56(4), Schneeweiss, G. M. (2016). Mechanistic model of evolutionary
1006–­1018. rate variation en route to a nonphotosynthetic lifestyle in plants.
Preuss, M., & Zuccarello, G. C. (2018). Three new red algal par- Proceedings of the National Academy of Sciences, 113(32),
asites from New Zealand: Cladhymenia oblongifoliaphila 9045–­9050.
sp. nov.(Rhodomelaceae), Phycodrys novae-­zelandiaephila Wicke, S., & Naumann, J. (2018). Molecular evolution of plastid
sp. nov.(Delesseriaceae) and Judithia parasitica sp. nov. genomes in parasitic flowering plants. In S. M. Chaw & R. K.
(Kallymeniaceae). Phycologia, 57(1), 9–­19. Jansen (Eds.), Advances in botanical research: Vol. 85. Plastid
Preuss, M., & Zuccarello, G. C. (2019). Development of the red algal genome evolution (pp. 315–­347). Academic Press.
parasite vertebrata aterrimophila sp. nov.(Rhodomelaceae, Wise, R. R. (2007). The diversity of plastid form and function. In R.
Ceramiales) from New Zealand. European Journal of R. Wise & J. K. Hoober (Eds.), The structure and function ofp-
Phycology, 54(2), 175–­183. Plastids (pp. 3–­26). Springer.
15298817, 2023, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/jpy.13373 by Faculty Of Medicine, Library Chulalongkorn University, Wiley Online Library on [18/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
962 |    PREUSS et al.

Yu, J., Wang, C., & Gong, X. (2017). Degeneration of photosyn-


thetic capacity in mixotrophic plants, iand Pyrola decorata Table S2. Sequence of custom designed primer
(Ericaceae). Plant Diversity, 39(2), 80–­88. sequences to check areas with frameshift mutations in
Zuccarello, G. C., & Lokhorst, G. M. (2005). Molecular phylogeny annotated plastid genomes in J. tasmanica.
of the genus Tribonema (Xanthophyceae) using rbcL gene se- Table S3. Accession numbers of plastid genomes,
quence data: Monophyly of morphologically simple algal spe-
cox1, and rbcL sequences used for phylogenetic
cies. Phycologia, 44(4), 384–­392.
Zuccarello, G. C., Moon, D., & Goff, L. J. (2004). A phylogenetic study analyses in alphabetical order by species. # indicates
of parasitic genera placed in the family Choreocolacaceae rbcL sequence was extracted from the complete plastid
(Rhodophyta). Journal of Phycology, 40(5), 937–­945. genome of the red alga.
Table S4. Key parameter settings used in phylogenetic
S U PP O R T I N G I N FO R M AT I O N analyses.
Additional supporting information can be found online Table S5. Summary of all genes and their length, in bp,
in the Supporting Information section at the end of this present in plastid genomes excluding unclassified open
article. reading frames of Janczewskia parasites and related
Figure S1. Annotated plastid genomes including are grouped alphabetically by protein coding genes,
protein coding genes and rRNAs (except for ORFs) rRNA, and tRNA. Lengths between species are in bold.
from Janczewskia tasmanica and one of its hosts
(Laurencia elata) and J. verruciformis and its host
(Laurencia catarinensis) and closest relative (Laurencia
obtusa). Host and parasite combinations are indicated How to cite this article: Preuss, M., Díaz-­Tapia,
by arrows. No gene arrangements, or gene loss, P., Verbruggen, H., & Zuccarello, G. C. (2023).
are found between any of the compared plastid Gene-­rich plastid genomes of two parasitic red
genomes. Pseudogenes are indicated with asterisks in algal species, Laurencia australis and L.
J. tasmanica. verruciformis (Rhodomelaceae, Ceramiales), and
Table S1. Location and GenBank Accession numbers a taxonomic revision of Janczewskia. Journal of
for all sequenced samples. # host species unknown Phycology, 59, 950–962. https://doi.org/10.1111/
and letter a-­e indicate the different host and parasite jpy.13373
combinations.

You might also like